首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The orientation of lipid headgroups may serve as a powerful sensor of electrostatic interactions in membranes. As shown previously by 2H NMR measurements, the headgroup of phosphatidylcholine (PC) behaves like an electrometer and varies its orientation according to the membrane surface charge. Here, we explored the use of solid-state 14N NMR as a relatively simple and label-free method to study the orientation of the PC headgroup in model membrane systems of varying composition. We found that 14N NMR is sufficiently sensitive to detect small changes in headgroup orientation upon introduction of positively and negatively charged lipids and we developed an approach to directly convert the 14N quadrupolar splittings into an average orientation of the PC polar headgroup. Our results show that inclusion of cholesterol or mixing of lipids with different length acyl chains does not significantly affect the orientation of the PC headgroup. In contrast, measurements with cationic (KALP), neutral (Ac-KALP), and pH-sensitive (HALP) transmembrane peptides show very systematic changes in headgroup orientation, depending on the amount of charge in the peptide side chains and on their precise localization at the interface, as modulated by varying the extent of hydrophobic peptide/lipid mismatch. Finally, our measurements suggest an unexpectedly strong preferential enrichment of the anionic lipid phosphatidylglycerol around the cationic KALP peptide in ternary mixtures with PC. We believe that these results are important for understanding protein/lipid interactions and that they may help parametrization of membrane properties in computational studies.  相似文献   

2.
Bacteriorhodopsin (BR), a membrane protein found in Halobacterium salinarum, functions as a light-driven proton pump. The Schiff base region has a quadrupolar structure with positive charges located at the protonated Schiff base and Arg82, and the counterbalancing negative charges located at Asp85 and Asp212. The quadropole inside the protein is stabilized by three water molecules, forming a roughly planar pentagonal cluster composed of these waters and two oxygens of Asp85 and Asp212 (one from each carboxylate side chain). It is known that BR lacks proton-pumping activity if Asp85 or Asp212 is neutralized by mutation, but binding of Cl- has different functional effects in mutants at these positions. Binding of Cl- to D85T converts into a chloride ion pump (Sasaki, J., Brown, L. S., Chon, Y.-S., Kandori, H., Maeda, A., Needleman, R., and Lanyi, J. K. (1995) Science 269, 73-75). On the other hand, photovoltage measurements suggested that binding of Cl- to D212N restores the proton-pumping activity at low pH (Moltke, S., Krebs, M. P., Mollaaghababa, R., Khorana, H. G., and Heyn, M. P. (1995) Biophys. J. 69, 2074-2083). In this paper, we studied halide-bound D212N mutant BR in detail. Light-induced pH changes in a suspension of proteoliposomes containing D212N(Cl-) at pH 5 clearly showed that Cl- restores the proton-pumping activity. Spectral blue-shift induced by halide binding to D212N indicates that halides affect the counterion of the protonated Schiff base, whereas much smaller halide dependence of the lambdamax than in D85T suggests that the binding site is distant from the chromophore. In fact, the K minus BR difference Fourier-transform infrared (FTIR) spectra of D212N at 77 K exhibit little halide dependence for vibrational bands of retinal and protein. The only halide-dependent bands were the C=N stretch of Arg82 and some water O-D stretches, suggesting that these groups constitute a halide-binding pocket. A strongly hydrogen-bonded water molecule is observed for halide-bound D212N, but not for halide-free D212N, which is consistent with our hypothesis that such a water molecule is a prerequisite for proton-pumping activity of rhodopsins. We concluded that halide binding near Arg82 in D212N restores the water-containing hydrogen-bonding network in the Schiff base region. In particular, the ion pair formed by the Schiff base and Asp85 through a strongly hydrogen-bonded water is essential for the proton-pumping activity of this mutant and may be controlled by the halide binding to the distant site.  相似文献   

3.
Methylation of 1,N6-ethenoadenosine (εAdo) gives a mixture of N1- and N9-quaternized methyl-3-β-D -ribofuranosylimidazo[2,1-i] purinium salts (m1εAdo+ and m9εAdo+, respectively). The ratio of the two forms of the protonated εAdo [H1εAdo+]/[H9εAdo+] has been estimated to be approximately 0.10 by comparing the uv absorption spectra of the protonated species of εAdo and the two nontautomerizable model compounds. In relation to a study on the protonation effect on the fluorescence of εAdo, we have now determined the effect of quaternization on the fluorescence spectra at 293 and 77 K. We have found that m1εAdo+ and m9εAdo+ are both fluorescent, and the high degree of coincidence between the fluorescence spectra of εAdo and m1εAdo+ at pH 7 is noted. The m1εAdo+ singlet form is a more efficient fluorescer than the m9εAdo+ ion at room temperature (quantum yields of 0.43 and 0.11, respectively). All the results which are presented in this paper are consistent with the picture that there exist more than one species responsible for the fluorescence of εAdo, depending on the environment of the molecule in aqueous solution (temperature and pH of solvent).  相似文献   

4.
T B Shin  R Leventis  J R Silvius 《Biochemistry》1991,30(30):7491-7497
Fluorescence spectroscopy has been used to monitor the partitioning of a series of exchangeable neutral phospholipid probes, labeled with carbazole, indolyl or diphenylhexatrienyl moieties, between large unilamellar vesicles containing 1-palmitoyl-2-oleoylphosphatidylcholine (POPC), 1,2-dioleoyloxy-3-(trimethylammonio) propane (DOTAP) or N-hexadecyl-N-(9-octadecenyl)-N,N-dimethylammonium chloride (HODMA). Phosphatidylethanolamine (PE) probes desorb from POPC-containing vesicles at markedly slower rates than do phosphatidylcholine (PC) probes with the same acyl chains. The rate of probe desorption from such vesicles is progressively enhanced by successive N-methylations of the amino group but not by methylation of C-2 of the ethanolamine moiety, a modification that leaves unaltered the hydrogen-bonding capacity of the polar headgroup. By contrast, the rates of desorption of different probes (with the same acyl chains) from HODMA or from DOTAP vesicles are much more comparable and reflect no clear systematic influence of the headgroup hydrogen-bonding capacity. Equilibrium-partitioning measurements indicate that the relative affinities of different probes for PC-rich vesicles, in competition with HODMA or DOTAP vesicles, increase with increasing hydrogen-bonding capacity of the probe headgroup in the order PC less than N,N-dimethyl PE less than N-methyl PE less than PE approximately phosphatidyl-2-amino-1-propanol. From such partitioning data, we estimate that interlipid hydrogen-bonding interactions (in competition with lipid-water interactions) contribute roughly -300 cal mol-1 to the free energy of a PE molecule in a hydrated liquid-crystalline phospholipid bilayer; this free-energy contribution is somewhat smaller, but still significant, for N-mono- and dimethylated PE's.  相似文献   

5.
Increasing methylation of the headgroup in DPPE results in an increase of minimum area per molecule in highly compressed monolayers at the air-water interface. The shape of solid domains, as observed by epifluorescence microscopy, also exhibits marked changes upon increasing headgroup methylation. Branching domains are observed in DPPE and DP(Me)PE, whereas U-shaped or round domains are observed in DP(Me)2PE and DPPC under our experimental conditions. The domain shape is determined more by the headgroup methylatin than by the corresponding shift in critical temperatures, as shown by the study of PCs of different acyl chain moieties. In mixed lipid monolayers, PC (phosphatidylcholine) and PE (phosphatidylethanolamine) do not mix ideally, as indicated by the non-linear variation of the average area per molecule with composition, and by distinct domain shapes in LE/LC (liquid expanded/liquid condensed) coexisting phases representing PE-enriched or PC-enriched domains in those mixed monolayers.  相似文献   

6.
The polar interface of membranes containing phosphatidylglycerol or cholesterol was studied by (2)H nuclear magnetic resonance (NMR) as a function of membrane hydration. The membranes were macroscopically aligned and hydrated with deuterium oxide. Water uptake and membrane annealing was achieved under NMR control, using a novel hydration technique. Well-resolved (2)H quadrupolar doublets were obtained from individual hydroxyl residues and from the interlamellar water. The response of the phosphatidylglycerol headgroup and of the cholesterol molecule to the spontaneous evaporation of interlamellar water could be thus monitored continuously. It is shown that the phosphatidylglycerol headgroup undergoes changes of conformation and average orientation with respect to the membrane surface and that the off-axis motion of the cholesterol molecule decreases. The deuteron exchange between hydroxyl residues and surface-associated D(2)O was determined by an inversion transfer technique. The exchange rates of the hydroxyl residues in the phosphatidylglycerol headgroup were different and depended strongly on the total hydration of the membrane. Significantly lower and almost hydration-independent rates were obtained for cholesterol. These results will be discussed with reference to earlier reports on the headgroup dynamics of phosphatidylglycerol and on the interaction of cholesterol with the membrane-water interface.  相似文献   

7.
Nucleosides dissolved in aqueous buffered solutions undergo ionization during direct introduction of the solution into a mass spectrometer using a thermospray interface. The principal ions formed represent the protonated molecule, the corresponding protonated free base, and sugar. In addition to potential utility for characterization of new nucleosides, the technique can be used to monitor nucleosides separated from enzymatic hydrolysates by liquid chromatography. The selectivity of chromatographic detection is significantly greater than with UV absorbance alone so that independent detection of components of unresolved chromatographic peaks is usually possible. Detection limits, with signal/noise greater than 10 for most nucleosides, are approximately 0.1-1 ng per component for selected ion monitoring and 10-50 ng for full-scan mass spectra. Examples are given from the detection of modified nucleosides in enzymatic hydrolysates of 0.05 A260 units (2.5 micrograms) of rabbit liver tRNAVal and of unfractionated H. volcanii tRNA.  相似文献   

8.
Monolayers of an enantiomeric and a racemic triple-chain phosphatidylcholine (PC) at the air/water interface are studied by film balance measurements and x-ray diffraction. Although the area per three tails exceeds that per head, we observe tail ordering dependent on headgroup chirality and chain tilt. This indicates lateral headgroup interactions. The influence of the chiral carbon is suppressed at higher lateral pressures, and a centered-rectangular unit cell with tails tilted into the nearest neighbor (NN) direction is observed for both the enantiomer and the racemate. The distortion of the lattice changes at medium pressures from NN to NNN (next-nearest neighbor direction) with decreasing temperature. The phase behavior of the racemate at 15 degrees C is compared with that of a triple-chain PC with a branched chain of reduced length. Whereas the PC with the longer branched chain exhibits only a NN tilted phase at all pressures, the PC with the shorter branched chain has a rich polymorphism (NNN-NN-upright hexagonal packing) under increased lateral pressure.  相似文献   

9.
10.
Liposomal formulation of curcumin is an important therapeutic agent for the treatment of various cancers. Despite extensive studies on the biological effects of this formulation in cancer treatment, much remains unknown about curcumin–liposome interactions. Understanding how different lipid bilayers respond to curcumin molecule may help us to design more effective liposomal curcumin. Here, we used molecular dynamics simulation method to investigate the behavior of curcumin in two lipid bilayers commonly used in preparation of liposomal curcumin, namely dipalmitoylphosphatidylcholine (DPPC) and dimyristoylphosphatidylglycerol (DMPG). First, the free energy barriers for translocation of one curcumin molecule from water to the lipid bilayer were determined by using the potential of mean force (PMF). The computed free energy profile exhibits a global minimum at the solvent–headgroup interface (LH region) for both lipid membranes. We also evaluated the free energy difference between the equilibrium position of curcumin in the lipid bilayer and bulk water as the excess chemical potential. Our results show that curcumin has the higher affinity in DMPG compared to DPPC lipid bilayer (?8.39 vs. ?1.69 kBT) and this is related to more hydrogen bond possibility for curcumin in DMPG lipid membrane. Next, using an unconstrained molecular dynamic simulation with curcumin initially positioned at the center of lipid bilayer, we studied various properties of each lipid bilayer system in the presence of curcumin molecule that was in full agreement with PMF and experimental data. The results of these simulation studies suggest that membrane composition could have a large effect on interaction of curcumin–lipid bilayer.  相似文献   

11.
An extensive 100-ns molecular dynamics simulation of lipid bilayer composed of mixture of phosphatidylethanolamine (PE) and phosphatidylglycerol (PG) was performed to elucidate the role of PGs to the stability of bacterial membranes. In addition, a control simulation of pure PE over 150 ns was performed. We observed that PGs decrease both the PE headgroup protrusions into the water phase, and the PE headgroup motion along bilayer normal. The above effects are caused by stronger inter-lipid interactions in the mixed bilayer: the number of hydrogen bonds created by PEs is 34% higher in the mixed than in the pure bilayer. Another contribution is due to the numerous ion-mediated inter-lipid links, which strongly enhance interface stability. That provides a plausible mechanism for preventing lipid desorption from the membrane, for example, under the influence of an organic solvent. A more compact and less dynamic interface structure also decreases membrane permeability. That provides a possible mechanism for stabilizing, e.g., bacterial membranes.  相似文献   

12.
The polar headgroup contribution to monolayer behavior of dipalmitoylglucosylglycerol has been examined through studies of 1,2-di-O-palmitoyl-3-O-(alpha-D-glucopyranosyl)-sn-glycerol (di-16:0-alpha GlcDG) and 1,2-di-O-palmitoyl-3-O-(beta-D-glucopyranosyl)-sn-glycerol (di-16:0-beta GlcDG) in which the sugar headgroup is linked via an alpha or beta linkage to the diacylglycerol moiety. The results indicate that the limiting areas per molecule of the resultant condensed states are smaller than those of the corresponding phosphatidylcholine (DPPC) but larger than those of dipalmitoylphosphatidylethanolmine (DPPE). In the expanded state, while the areas per molecule are similar to those of DPPC at low pressures, both glycolipids occupy smaller areas at higher pressures. The expanded-state areas of the glucolipids are also slightly greater than those of DPPE. The initial compressional phase transition pressure of the glucolipid liquid-expanded/liquid-condensed transition (pi t) is, however, less sensitive to temperature than are the pi t values of phospholipids. Both of these effects must relate to strong headgroup/water interactions, which, in turn, result in a stabilization of the liquid-expanded states. In the expanded states the alpha anomers are slightly less tightly packed than the beta anomers, as is indicated by the somewhat higher areas per molecule of the expanded states and the lower transition temperatures. These differences in chain-melting temperatures are slightly smaller than those observed in bilayers. While the areas per molecule of the dipalmitoyl glucolipids are greater than those of dipalmitoylphosphatidylethanolamine, they nevertheless exhibit a greater tendency to form nonbilayer structures. Such observations indicate that other factors besides geometric shape play a role in bilayer/nonbilayer transitions.  相似文献   

13.
C H Hsieh  S C Sue  P C Lyu    W G Wu 《Biophysical journal》1997,73(2):870-877
Diphytanoylphosphatidylcholine (DPhPC) has often been used in the study of protein-lipid interaction and membrane channel activity, because of the general belief that it has high bilayer stability, low ion leakage, and fatty acyl packing comparable to that of phospholipid bilayers in the liquid-crystalline state. In this solid-state 31P and 2H NMR study, we find that the membrane packing geometry and headgroup orientation of DPhPC are highly sensitive to the temperature studied and its water content. The phosphocholine headgroup of DPhPC starts to change its orientation at a water content as high as approximately 16 water molecules per lipid, as evidenced by hydration-dependent 2H NMR study at room temperature. In addition, a temperature-induced structural transition in the headgroup orientation is detected in the temperature range of approximately 20-60 degrees C for lipids with approximately 8-11 water molecules per DPhPC. Dehydration of the lipid by one more water molecule leads to a nonlamellar, presumably cubic, phase formation. The lipid packing becomes a hexagonal phase at approximately 6 water molecules per lipid. A phase diagram of DPhPC in the temperature range of -40 degrees C to 80 degrees C is thus constructed on the basis of NMR results. The newly observed hydration-dependent DPhPC lipid polymorphism emphasizes the importance of molecular packing in the headgroup region in modulating membrane structure and protein-induced pore formation of the DPhPC bilayer.  相似文献   

14.
The binding of trimethoprim and [1,3,2-amino-15N3]-trimethoprim to Lactobacillus casei dihydrofolate reductase has been studied by 15N and 1H NMR spectroscopy. 15N NMR spectra of the bound drug were obtained by using polarisation transfer pulse sequences. The 15N chemical shifts and 1H-15N spin-coupling constants show unambiguously that the drug is protonated on N1 when bound to the enzyme.The N1-proton resonance in the complex has been assigned using the 15N-enriched molecule. The temperature-dependence of the linewidth of this resonance has been used to estimate the rate of exchange of this proton with the solvent: 160±10s-1 at 313 K, with an activation energy of 75 (±9) kJ·mole-1. This is considerably faster than the dissociation rate of the drug from this complex, demonstrating that there are local fluctuations in the structure of the complex.  相似文献   

15.
To better understand bilayer property dependency on lipid electrostatics and headgroup size, we use atomistic molecular dynamics simulations to study negatively charged and neutral lipid membranes. We compare the negatively charged phosphatidic acid (PA), which at physiological pH and salt concentration has a negative spontaneous curvature, with the negatively charged phosphatidylglycerol (PG) and neutrally charged phosphatidylcholine (PC), both of which have zero spontaneous curvature. The PA lipids are simulated using two different sets of partial charges for the headgroup and the varied charge distribution between the two PA systems results in significantly different locations for the Na+ ions relative to the water/membrane interface. For one PA system, the Na+ ions are localized around the phosphate group. In the second PA system, the Na+ ions are located near the ester carbonyl atoms, which coincides with the preferred location site for the PG Na+ ions. We find that the Na+ ion location has a larger effect on bilayer fluidity properties than lipid headgroup size, where the Alipid and acyl chain order parameter values are more similar between the PA and PG bilayers that have Na+ ions located near the ester groups than between the two PA bilayers.  相似文献   

16.
In this study, the roles of two anionic phospholipids—phosphatidic acid (PA), which is an important signaling molecule, and cardiolipin (CL), which plays a crucial role in the bioenergetics of the cell—in stabilizing the oligomeric structure of potassium channel KcsA were determined. The stability of KcsA was drastically increased as a function of PA or CL content (mol%) in phosphatidylcholine (PC) bilayers. Deletion of the membrane-associated N terminus significantly reduced channel stability at high levels of PA content; however, the intrinsic stability of this protein was marginally affected in the presence of CL. These studies indicate that the electrostatic-hydrogen bond switch between PA and N terminus, involving basic residues, is much stronger than the stabilizing effect of CL. Furthermore, the unique properties of the PA headgroup alter protein assembly and folding properties differently from the CL headgroup, and both lipids stabilize the tetrameric assembly via their specific interaction on the extra- or the intracellular side of KcsA.  相似文献   

17.
Binding of thiocyanate and cyanide ions to Mn(III) protoporphyrin-apohorseradish peroxidase complex [Mn(III)HRP] was investigated by relaxation rate measurements (at 50.68 MHz) of 15N resonance of SC15N- and C15N-. At pH = 4.0 the apparent dissociation constant (KD) for thiocyanate and cyanide binding to Mn(III)HRP was deduced to be 156 and 42 mM, respectively. The pH dependence of the 15N line width as well as apparent dissociation constant for thiocyanate and cyanide binding were quantitatively analyzed on the basis of a reaction scheme in which thiocyanate and cyanide in deprotonated form bind to the enzyme in a protonated form. The binding of thiocyanate and cyanide to Mn(III)HRP was found to be facilitated by protonation of an ionizable group on the enzyme [Mn(III)HRP] with a pKa = 4.0. From competitive binding studies it was shown that iodide, thiocyanate and cyanide bind to Mn(III)HRP at the same site; however, the binding site for resorcinol is different. The apparent dissociation constant for iodide binding deduced from competitive binding studies was found to be 117 mM, which agrees very well with the iodide binding to ferric HRP. The binding of thiocyanate and cyanide was shown to be away from the metal center and the distance of the 15N of thiocyanate and cyanide from the paramagnetic manganese ion in Mn(III)HRP was found to be 6.9 and 6.6 A, respectively. Except for cyanide binding, these observations parallel with the iodide and thiocyanate ion binding to native Fe(III)HRP. Water proton relaxivity measurements showed the presence of a coordinated water molecule to Mn(III)HRP with the distance of Mn-H2O being calculated to be 2.6 A. The slow reactivity of H2O2 towards Mn(III)HRP could be attributed to the presence of water at the sixth coordination position of the manganese ion.  相似文献   

18.
By varying the pH, the D85N mutant of bacteriorhodopsin provides models for several photocycle intermediates of the wild-type protein in which D85 is protonated. At pH 10.8, NMR spectra of [zeta-(15)N]lys-, [12-(13)C]retinal-, and [14,15-(13)C]retinal-labeled D85N samples indicate a deprotonated, 13-cis,15-anti chromophore. On the other hand, at neutral pH, the NMR spectra of D85N show a mixture of protonated Schiff base species similar to that seen in the wild-type protein at low pH, and more complex than the two-state mixture of 13-cis,15-syn, and all-trans isomers found in the dark-adapted wild-type protein. These results lead to several conclusions. First, the reversible titration of order in the D85N chromophore indicates that electrostatic interactions have a major influence on events in the active site. More specifically, whereas a straight chromophore is preferred when the Schiff base and residue 85 are oppositely charged, a bent chromophore is found when both the Schiff base and residue 85 are electrically neutral, even in the dark. Thus a "bent" binding pocket is formed without photoisomerization of the chromophore. On the other hand, when photoisomerization from the straight all-trans,15-anti configuration to the bent 13-cis,15-anti does occur, reciprocal thermodynamic linkage dictates that neutralization of the SB and D85 (by proton transfer from the former to the latter) will result. Second, the similarity between the chromophore chemical shifts in D85N at alkaline pH and those found previously in the M(n) intermediate of the wild-type protein indicate that the latter has a thoroughly relaxed chromophore like the subsequent N intermediate. By comparison, indications of L-like distortion are found for the chromophore of the M(o) state. Thus, chromophore strain is released in the M(o)-->M(n) transition, probably coincident with, and perhaps instrumental to, the change in the connectivity of the Schiff base from the extracellular side of the membrane to the cytoplasmic side. Because the nitrogen chemical shifts of the Schiff base indicate interaction with a hydrogen-bond donor in both M states, it is possible that a water molecule travels with the Schiff base as it switches connectivity. If so, the protein is acting as an inward-driven hydroxyl pump (analogous to halorhodopsin) rather than an outward-driven proton pump. Third, the presence of a significant C [double bond] N syn component in D85N at neutral pH suggests that rapid deprotonation of D85 is necessary at the end of the wild-type photocycle to avoid the generation of nonfunctional C [double bond] N syn species.  相似文献   

19.
Monolayer studies at the air/water interface were carried out on the major tetraether (caldarchaeol-) derived phosphoglycolipid, Glcp-alpha(1-2)-Galf-beta(1-1)-caldarchaeol-phosphoglycerol (PGC-I), the major diether (archaeol-) derived glycolipid, Glcp-alpha(1-2)-Galf-beta(1-1)-archaeol (DGA-I), the major archaeol-derived phospholipids, phosphatidyl-N,N dimethylaminopentanetetrol (PPDAA) and phosphatidyl-N,N,N-trimethylaminopentanetetrol (PPTAA) and the minor caldarchaeol-derived glycolipid, Glcp-alpha(1-2)-Galf-beta(1-1)-caldarchaeol (DGC-I) isolated from the methanogenic archaebacterium, Methanospirillum hungatei. The compression isotherms obtained showed that the two tetraether lipids had molecular surface areas about twice those of the diether lipids at all surface pressures, suggesting that both polar headgroups of the tetraether lipids are anchored into the aqueous subphase, even at the collapse pressure pi c. A U-shaped hydrocarbon chain conformation thus appears to be preferred for the tetraether lipids at the air/water interface, rather than an extended chain arrangement. The compression isotherms of the two tetraether lipids PGC-I and DGC-I were very similar at pH 0, both molecules being uncharged, but at pH 5.6 or 8, PGC-I films were much more expanded than the neutral DGC-I, due to ionization of the phosphate group in PGC-I and the resulting charge-charge repulsion. Monolayers of the zwitterionic diether phospholipids PPDAA and PPTAA were much less compressible than the glycosylated lipids, PGC-I, DGC-I and DGA-I, because the latter lipids contain the more compressible diglycosyl headgroup, oriented in horizontal conformation at low surface pressures, compared to the lower compressibility of the zwitterionic headgroup in the vertical conformation, particularly at pH 0 and 5.6.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
The new compound trimethylene-N(6),N(6')-bisadenine (L), in which two adenine molecules are linked together by a trimethylene bridge that connects the N(6) atoms, has been prepared. Reaction of L with HgCl(2) and ZnCl(2) in concentrated HCl solution leads to crystalline solids. The X-ray characterisation of the Hg(II) complex (H(2)L)[HgCl(4)].3H(2)O reveals that it is an outer-sphere complex in which the ligand is protonated at N(1) and N(1'). In contrast, the structure of the complex [H(2)L(ZnCl(3))(2)].2H(2)O shows the ligand co-ordinated to two different Zn(II) ions through the N(7) of both adenine fragments, the protons being located on the N(1) atoms. The latter compound constitutes the first crystallographic evidence of an inner sphere complex with bis-adenines and, for this reason, an equilibrium study was carried out on the Zn(II)-L-H(+) system. Potentiometric studies indicate that L is protonated in aqueous solution to form HL(+) and H(2)L(2+) with logK(H) values of 4.42 and 3.35 (25 degrees C, 0.10 M KNO(3)). The data from potentiometric titrations in the presence of Zn(2+) can be analysed considering the formation of the species LZn(2+), HLZn(3+), LZn(2)(4+) and HLZn(2)(5+), whose stability constants exceed the value expected for a monodentate interaction of the metal ion with adenine and suggest the possibility of a polydentate behaviour of L in the pH range 2.5-5.0. In contrast, spectrophotometric titrations carried out under conditions similar to those used in the synthetic work (1 M HCl) can be fitted with a model involving exclusively the H(2)LZn(4+) and H(2)LZn(2)(6+) species with logK(M) values reasonable for the interaction of Zn(II) with the N(7) of the protonated adenine fragments. Despite the H(2)LZn(2)(6+) species has a low stability, the spectrophotometric results are in agreement with its formation under the conditions in which the solid complex was prepared.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号